...

State shift in Deccan volcanism at the Cretaceous

by user

on
Category: Documents
23

views

Report

Comments

Transcript

State shift in Deccan volcanism at the Cretaceous
R ES E A RC H | R E PO R TS
EARTH HISTORY
State shift in Deccan volcanism at
the Cretaceous-Paleogene boundary,
possibly induced by impact
Paul R. Renne,1,2* Courtney J. Sprain,1,2 Mark A. Richards,2 Stephen Self,2
Loÿc Vanderkluysen,3 Kanchan Pande4
Bolide impact and flood volcanism compete as leading candidates for the cause of
terminal-Cretaceous mass extinctions. High-precision 40Ar/39Ar data indicate that these
two mechanisms may be genetically related, and neither can be considered in isolation.
The existing Deccan Traps magmatic system underwent a state shift approximately
coincident with the Chicxulub impact and the terminal-Cretaceous mass extinctions, after
which ~70% of the Traps' total volume was extruded in more massive and more episodic
eruptions. Initiation of this new regime occurred within ~50,000 years of the impact,
which is consistent with transient effects of impact-induced seismic energy. Postextinction
recovery of marine ecosystems was probably suppressed until after the accelerated
volcanism waned.
T
he Deccan Traps are the most recent of
several large (>106 km3) continental flood
basalt provinces that are circumstantially
implicated in mass extinctions (1, 2). The
extent to which Deccan volcanism was a
factor in the biotic crises that terminated the
Mesozoic Era has been heavily debated, largely
because the Chicxulub bolide impact provides a
plausible mechanism for severe and abrupt environmental perturbations (3). The temporal coincidence between the impact and mass extinctions
at the Cretaceous-Paleogene boundary (KPB) is
well established (4) and implicates the bolide
impact as a forcing mechanism. However, the
possibility of a contributing role for volcanism,
presumably through the discharge of climatemodifying gases, remains plausible in view of data
showing that the immense flood basalt eruptions
of the Deccan Traps spanned the KPB. A temporal
correlation between other major flood volcanic
events and extensive environmental transitions
(1, 2) illustrates the potential for the Deccan Traps
alone to have caused the KPB extinctions. We
combined high-precision 40Ar/39Ar dating of the
lavas with previously reported U/Pb dates (5)
and lava volume estimates (6) to infer that the
Chicxulub impact initiated a substantial acceleration of Deccan volcanism within ~50 thousand
years (ky). This probably contributed to the extinctions and moderated subsequent recovery.
The general coincidence of the Deccan Traps
with the KPB, evident in geochronologic and paleomagnetic data (7), is not established with sufficient precision or stratigraphic control to locate
1
Berkeley Geochronology Center, 2455 Ridge Road, Berkeley,
CA 94709, USA. 2Department of Earth and Planetary
Science, University of California–Berkeley, Berkeley, CA
94720, USA. 3Department of Biodiversity, Earth and
Environmental Science, Drexel University, Philadelphia, PA
19104, USA. 4Department of Earth Sciences, Indian Institute
of Technology Bombay, Powai, Mumbai 400 076, India.
*Corresponding author. E-mail: [email protected]
76
the boundary within the lava stratigraphy. Limited constraints on the detailed history and tempo
of volcanism obscure possible mechanisms that
would relate the Deccan Traps to KPB phenomena, because these factors modulate volatile input
to the atmosphere (8). The absence of reliable
estimates of magma volume over time challenges
models that suggest extrusion of the Traps in
three discrete pulses (9, 10). The geochemically
defined stratigraphy of Deccan Group volcanics
in the best-studied region, the Western Ghats, is
divided into formations composing the Kalsubai,
Lonavala, and Wai Subgroups, in ascending order. Each formation comprises multiple eruptive
units. Paleomagnetic secular variation suggests
that the eruptions were highly episodic, with a
lower mean eruption frequency in the youngest
and most widely distributed subgroup, the Wai
(11, 12). Lavas with broadly contemporaneous
ages and geochemical and isotopic affinities
with the Wai Subgroup occur as far north as
Rajasthan (13), as far northeast as the Mandla
lobe (~1000 km from the Western Ghats escarpment) (14, 15), and as far east as Rajahmundry on
the Bay of Bengal coast (16).
The Wai Subgroup contains numerous oxidized
horizons (“red boles”) between lava flows, interpreted as paleosols (17–19), which are consistent
with the inference of a lower eruption frequency.
Red boles are much less prevalent in the underlying Lonavala and Kalsubai Subgroups (11, 20),
providing possible evidence for variable long-term
eruption rates during the lifetime of the Deccan
magma system.
A discontinuity in lava geochemistry, flow volumes, and feeder dike orientations occurs between the Lonavala and Wai Subgroups (6). This
discontinuity coincides with geomorphic evidence
of widespread fractures and an abrupt change in
susceptibility to erosion, approximately at the contact between these subgroups (6). It also coincides
with the transition to more episodic, albeit more
voluminous, eruptions in the Wai Subgroup, as
inferred from paleomagnetic secular variation and
the frequency of red bole horizons. Existing age
constraints are compatible with the hypothesis
that the KPB coincides with the transition from
the Lonavala Subgroup (Bushe Formation) to the
Wai Subgroup (Poladpur Formation) (6).
Fig. 1. Stratigraphic and geographic context of the samples. (A) Stratigraphic sections with dated horizons indicated, corresponding to the five traverses shown by red lines in (B). Ages (millions of years ago)
are shown with uncertainties (SEM). Sample numbers for the dated horizons are given in parentheses.
(B) Stratigraphy adapted from (11) for traverses KAS (Shahapur-Igatpuri section), MAT (Matheran-Neral
section), BOR (Khopali-Khandala section), and AMB (Mahabaleshwar-Poladpur section). Stratigraphic data
for the MSJ traverse (Malshej Ghat section) were acquired in this study. The dotted black line connects the
bases of the traverses. The black rectangle in (C) shows the location of the area shown in (B).
EMBARGOED UNTIL 2PM U.S. EASTERN TIME ON THE THURSDAY BEFORE THIS DATE:
2 OCTOBER 2015 • VOL 350 ISSUE 6256
sciencemag.org SCIENCE
RE S EAR CH | R E P O R T S
Available geochronologic data are not sufficiently precise to resolve age variations of less than
100 ky, with the exception of U/Pb zircon data (5).
These data reveal protracted zircon age distributions, requiring subjective interpretation (7). Additionally, material suitable for U/Pb geochronology
is reported only in rare bodies interpreted as magma segregation features and in red boles, which
are sparse in both the Lonavala and Kalsubai
Subgroups. These limitations restrict both the extent and the overall resolution of Deccan stratigraphy that can be determined from the U/Pb
technique. Here, we report high-resolution 40Ar/39Ar
dating of igneous plagioclase, which has an unambiguous genetic relationship to the host lavas
and negligible retention of pre-eruptive radiogenic
40
Ar at basalt eruption temperatures >1100°C.
We used detailed stepwise heating, dense bracketing of samples with standards during neutron
irradiation, and detailed characterization of the
interfering nuclear reactions to obtain data of
appropriate precision and accuracy to clarify the
eruptive history of the Deccan Traps (7).
We obtained 40Ar/39Ar plateau ages (fig. S3)
for one to four aliquots of samples from each
formation in the Kalsubai Subgroup and from
the Ambenali Formation of the Wai Subgroup, in
multiple sections across the Western Ghats region
(Fig. 1). Where formation assignments were ambiguous, we analyzed samples geochemically to
confirm their placement in the chemical stratigraphy by which the formations are defined (21).
Weighted mean plateau ages placed in a composite stratigraphic section (Fig. 2) are generally
consistent with previously reported U/Pb zircon
data (5), within uncertainties.
Based on a composite stratigraphic section (6),
our data indicate rapid eruption of no less than
~70% of the Kalsubai Subgroup before the KPB,
over an interval of 173 ± 84 ky, as defined by the age
difference between our samples from the Jawhar
and Bhimashankar Formations (Fig. 2). The thicknesses of individual Kalsubai and Lonavala Subgroup formations vary substantially throughout the
province; thus, inferences about mean magma
generation rates based on values from any one
section may be misleading. More meaningful inferences can be drawn from best estimates of volume
(weighted by subgroup areal extent) for each formation (6), which indicate a volume of 71 × 103 km3
for this time interval, corresponding to a mean
magma generation rate of ~0.4 ± 0.2 km3/year.
The KPB, dated at 66.043 ± 0.010 (without
systematic sources) or ± 0.043 [with systematic
sources (7)] million years ago (22), occurred less
than 165 ± 68 ky after the emplacement of the
Kalsubai Subgroup, within the time represented
by the Khandala, Bushe, Poladpur, or possibly
lower Ambenali Formations. The time interval
between the eruption of the basal Bhimashankar
and uppermost Ambenali Formations is 547 ±
241 ky, during which ~300 × 103 km3 of lava was
extruded at a mean rate of ~0.6 km3/year. We
determined a magma generation rate for the Wai
Subgroup from U/Pb zircon dates (5) in the upper
Ambenali and mid-Mahabaleshwar Formations,
finding that a volume of 111 × 103 km3 of lava was
extruded at a mean rate of 0.9 ± 0.3 km3/year. The
mean long-term eruption rate represented by the
Wai Subgroup, where it is well constrained,
appears to be approximately double the rate represented by the Kalsubai Subgroup.
The increased mean magma production rate
during the Wai Subgroup eruptions coincides
with an abundance of red bole (weathering) horizons (Fig. 2), which first appear frequently in
the Ambenali Formation and persist into the
Mahabaleshwar Formation. The cumulative time
represented by red boles in the Western Ghats
has been estimated to be on the order of 300 ky
(11), most of which falls in the time interval represented by the Wai Subgroup. During the Wai
Subgroup time interval, it appears that the mean
eruption frequency decreased dramatically, whereas
the lava volume per flow field (per single eruptive
event) increased, so that the mean magma eruption rate approximately doubled. Alternatively,
the increased abundance of red boles in the Wai
Subgroup may reflect an increase in weathering
rates due to climate change (greenhouse conditions) after the KPB.
The decrease in eruption frequency coincides
generally with a transition from lava fields dominated by compound pahoehoe and rare a’a flows
in the Kalsubai Subgroup (23) to lava fields dominated by thicker, laterally extensive, inflated pahoehoe sheet lobes in the Wai Subgroup (20, 24).
The initiation of this transition appears to occur
across the Lonavala Subgroup, with compound
units absent in parts of the Khandala Formation
but present in the Bushe Formation. In our study
areas, sheet lobes begin to dominate near the base
of the Wai Subgroup within the Poladpur Formation, where individual sheet lobes reach the
greatest observed thickness of more than 60 m
(20). The sparse occurrence of red boles in the
Poladpur Formation suggests that the high eruption frequency continued until the time of the
Ambenali Formation, at which point large eruptions became punctuated by longer repose intervals (represented by the red boles). During
these repose intervals, occasional zircon-bearing,
distally sourced silicic tephras apparently accumulated in developing paleosols (5).
Fig. 2. Eruptive history of the
Western Ghats region.
(A) Composite stratigraphic
thicknesses and (B) cumulative
volumes, after (6), versus age
for formations in the Western
Ghats region. Included are
(i) all 40Ar/39Ar plagioclase dates
with uncertainties (SEM) of
300 ky or less, comprising data
from this study (filled circles)
and a previous study (open circles) (35), and (ii) previously
reported U/Pb zircon dates (red
triangles) (5). All ages are shown
with 1s uncertainties (bars),
including from systematic
sources. The apices of the triangles point in the direction of true
age in the event of a possible
bias in U/Pb ages (supplementary text). Sample positions are
scaled from formation thicknesses in the sections sampled, as shown in Fig. 1A. Red boles are scaled from previous studies (11, 20) and our observations. New
40
Ar/39Ar dates define the mean eruption rate represented by the Kalsubai Subgroup (0.4 T 0.2 km3/year); previously reported U/Pb zircon dates (5) define the
mean eruption rate represented by part of the Wai Subgroup (0.9 T 0.3 km3/year). The dotted red curve shows a parabolic fit, representing possible continuous
change in eruption rates between the time intervals of the Kalsubai and upper Wai Subgroups. The Panhala and Desur Formations, which overlie the
Mahabaleshwar Formation with limited extents, are not shown. Ma, millions of years ago.
EMBARGOED UNTIL 2PM U.S. EASTERN TIME ON THE THURSDAY BEFORE THIS DATE:
SCIENCE sciencemag.org
2 OCTOBER 2015 • VOL 350 ISSUE 6256
77
R ES E A RC H | R E PO R TS
The transition from high-frequency, low-volume
eruptions to low-frequency, high-volume eruptions
suggests a fundamental change in the magma
plumbing system. A state shift in the threshold
properties of one or more magma chambers would
have been required for the latter style of eruptions to occur. The Wai Subgroup lavas mark a
sharp increase in mantle contributions relative
to crustal contributions, are the least crustally contaminated of the entire Deccan Group (21, 25–27),
and overlie Bushe Formation lavas that are the
most crustally contaminated of the Deccan Group
(21, 25, 26). This suggests a reduction in the
magma-crust interface area, relative to magma
volume, through an expansion of a deep-crustal
magma chamber or a consolidation of multiple
chambers; a shorter residence time of magma in
the crust; or some combination of these factors.
A magma chamber expansion commencing
with the emplacement of the Poladpur Formation, coupled with longer repose times indicated
by the increased occurrence of red boles, would
be associated with more extensive crystal fractionation (28). Wai Subgroup lavas are the most highly fractionated in the entire sequence (21, 25, 27).
These and other independent lines of geochemical and tectonic evidence are all consistent with
a sudden increase in magma flux from the Deccan
mantle source (6). Larger flux rates would lead to
larger magma chambers (presumably at Moho
depths), longer time intervals to generate sufficient
buoyancy to drive eruptions (via some combination of crystal fractionation and volatile exsolution), and hence larger and less frequent eruptions
(29). We conclude that the Deccan magmatic system, at the source of the Western Ghats flood basalts and, by inference, the main portion of the
Deccan Volcanic Province, underwent a fundamental transition upon initiation of the Wai Subgroup eruptions.
Straightforward interpolation of the age-volume
data (Fig. 2B) suggests that this transition occurred within 50 ky of the Chicxulub impact and
the KPB. This is an even closer temporal coincidence than indicated by previous analyses (6),
which also suggested that strong seismic waves
produced by the impact could have triggered
increased volcanism. The close temporal coincidence of the impact and the accelerated volcanism
makes it difficult to deconvolve the environmental
perturbations attributable to each mechanism.
The KPB extinctions probably resulted from the
superposed effects of both phenomena.
The coincidence of widespread seismites at the
Triassic-Jurassic boundary (30), correlated with
the Central Atlantic Magmatic Province (31), may
represent another example of seismic triggering
of continental flood volcanism. The mechanisms
by which seismic events trigger volcanic activity
are not understood in detail (32), but they probably involve either a transient increase in the
effective permeability of existing volcanic systems or, perhaps, induced volatile exsolution
from supersaturated magma (33). These transient
effects may not account for the subsequent episodic repetition of larger-volume, lower-frequency
eruptions throughout the Wai Subgroup interval;
78
however, the repetition of such eruptions might
reflect a state change toward longer recharge times
due to magma chamber enlargement, perhaps
combined with other episodic seismic disturbances
(such as from large regional tectonic earthquakes).
Eruptions that emplaced the volumetrically dominant formations of the Wai Subgroup continued
for ~500 ky after the KPB, which is comparable
with the time lag between the KPB and the initial
stage of ecological recovery in marine ecosystems
(34). The effects of the Wai Subgroup eruptions
may have suppressed postextinction recovery until the final stages of the time interval represented
by the Mahabaleshwar Formation.
RE FERENCES AND NOTES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
V. E. Courtillot, P. R. Renne, C. R. Geosci. 335, 113–140 (2003).
P. B. Wignall, Earth Sci. Rev. 53, 1–33 (2001).
P. Schulte et al., Science 327, 1214–1218 (2010).
P. R. Renne et al., Science 339, 684–687 (2013).
B. Schoene et al., Science 347, 182–184 (2015).
M. A. Richards et al., Geol. Soc. Am. Bull. (2015).
Materials and methods are available as supplementary
materials on Science Online.
S. Self, M. Widdowson, T. Thordarson, A. E. Jay, Earth Planet.
Sci. Lett. 248, 518–532 (2006).
A. L. Chenet, X. Quidelleur, F. Fluteau, V. Courtillot, S. Bajpai,
Earth Planet. Sci. Lett. 263, 1–15 (2007).
G. Keller et al., Earth Planet. Sci. Lett. 341–344, 211–221 (2012).
A. L. Chenet et al., J. Geophys. Res. Solid Earth 114, B06103 (2009).
A. L. Chenet, F. Fluteau, V. Courtillot, M. Gerard, K. V. Subbarao,
J. Geophys. Res. Solid Earth 113, B04101 (2008).
A. Sen et al., J. Asian Earth Sci. 59, 127–140 (2012).
J. P. Shrivastava, R. A. Duncan, M. Kashyap, Lithos 224-225,
214–224 (2015).
Z. X. Peng, J. J. Mahoney, P. R. Hooper, J. D. Macdougall,
P. Krishnamurthy, J. Geophys. Res. Solid Earth 103, 29843 (1998).
S. Self, A. E. Jay, M. Widdowson, L. P. Keszthelyi, J. Volcanol.
Geotherm. Res. 172, 3–19 (2008).
P. Ghosh, M. R. G. Sayeed, R. Islam, S. M. Hundekari,
Palaeogeogr. Palaeoclimatol. Palaeoecol. 242, 90–109 (2006).
J. P. Shrivastava, M. Ahmad, S. Srivastava, J. Geol. Soc. India
80, 177–188 (2012).
M. Widdowson, J. N. Walsh, K. V. Subbarao, Geol. Soc. Spec.
Publ. 120, 269–281 (1997).
20. A. E. Jay, thesis, The Open University, Milton Keynes, UK (2005).
21. J. E. Beane, C. A. Turner, P. R. Hooper, K. V. Subbarao,
J. N. Walsh, Bull. Volcanol. 48, 61–83 (1986).
22. C. J. Sprain, P. R. Renne, G. P. Wilson, W. A. Clemens, Geol.
Soc. Am. Bull. 127, 393–409 (2015).
23. R. J. Brown, S. Blake, N. R. Bondre, V. M. Phadnis, S. Self, Bull.
Volcanol. 73, 737–752 (2011).
24. N. R. Bondre, R. A. Duraiswami, G. Dole, Bull. Volcanol. 66,
29–45 (2004).
25. P. C. Lightfoot, C. J. Hawkesworth, C. W. Devey, N. W. Rogers,
P. W. C. Van Calsteren, J. Petrol. 31, 1165–1200 (1990).
26. Z. X. Peng, J. J. Mahoney, P. Hooper, C. Harris, J. E. Beane,
Geochim. Cosmochim. Acta 58, 267–288 (1994).
27. J. J. Mahoney et al., Earth Planet. Sci. Lett. 60, 47–60 (1982).
28. C. J. Hawkesworth et al., J. Petrol. 41, 991–1006 (2000).
29. L. Karlstrom, M. Richards, J. Geophys. Res. Solid Earth 116,
B08216 (2011).
30. M. J. Simms, Geology 31, 557 (2003).
31. A. Marzoli et al., Science 284, 616–618 (1999).
32. M. Manga, E. Brodsky, Annu. Rev. Earth Planet. Sci. 34,
263–291 (2006).
33. R. J. Carey et al., J. Geophys. Res. Solid Earth 117, B11202 (2012).
34. H. K. Coxall, S. D'Hondt, J. C. Zachos, Geology 34, 297
(2006).
35. P. Hooper, M. Widdowson, S. Kelley, Geology 38, 839–842
(2010).
AC KNOWL ED GME NTS
This work was funded by the Ann and Gordon Getty Foundation
and by the Esper S. Larsen Fund of the University of California–
Berkeley. C.J.S. was supported by a NSF Graduate Research
Fellowship. L.V.’s participation was supported by NSF grant EAR1250440. We thank W. Alvarez, S. Finnegan, C. Guns, R. Ickert,
M. Manga, C. Marshall, A. Marzoli, R. Mundil, and H. Sheth for
discussion; T. Becker and A. Jaouni for laboratory assistance; and
H. Sheth for field assistance. Data are available in the
supplementary materials.
SUPPLEMENTARY MATERIALS
www.sciencemag.org/content/350/6256/76/suppl/DC1
Materials and Methods
Supplementary Text
Figs. S1 and S2
Table S1
References (36–69)
Database S1
8 June 2015; accepted 26 August 2015
10.1126/science.aac7549
WATER STRUCTURE
Ultrafast 2D IR spectroscopy of the
excess proton in liquid water
Martin Thämer,1 Luigi De Marco,1,2 Krupa Ramasesha,2*
Aritra Mandal,1,2 Andrei Tokmakoff 1†
Despite decades of study, the structures adopted to accommodate an excess proton in
water and the mechanism by which they interconvert remain elusive. We used ultrafast twodimensional infrared (2D IR) spectroscopy to investigate protons in aqueous hydrochloric
acid solutions. By exciting O–H stretching vibrations and detecting the spectral response
throughout the mid-IR region, we observed the interaction between the stretching and
bending vibrations characteristic of the flanking waters of the Zundel complex, [H(H2O)2]+,
at 3200 and 1760 cm−1, respectively. From time-dependent shifts of the stretch-bend cross
peak, we determined a lower limit on the lifetime of this complex of 480 femtoseconds.
These results suggest a key role for the Zundel complex in aqueous proton transfer.
A
cid-base chemistry and most biological redox
chemistry are governed by the transport of
protons through water. Aqueous proton
transfer is generally accepted to occur along
hydrogen bonds through sequential hops of
an excess proton from one solvating water mol-
ecule to the next. Although this widely accepted
picture, known as the Grotthuss mechanism,
captures the concept of long-range charge translocation without transport of a particular proton,
numerous basic questions remain regarding the
rapidly evolving structure of an aqueous proton (1).
EMBARGOED UNTIL 2PM U.S. EASTERN TIME ON THE THURSDAY BEFORE THIS DATE:
2 OCTOBER 2015 • VOL 350 ISSUE 6256
sciencemag.org SCIENCE
Fly UP